Base-Catalyzed Stereospecific Isomerization of Electron-Deficient

Sep 16, 2016 - 2,2-Diiododimedone: a mild electrophilic iodinating agent for the selective synthesis of α-iodoketones from allylic alcohols. Samuel ...
0 downloads 3 Views 2MB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Article

Base-Catalyzed Stereospecific Isomerization of ElectronDeficient Allylic Alcohols and Ethers through Ion-Pairing Samuel Martinez-Erro, Amparo Sanz-Marco, Antonio Bermejo Gómez, Ana Vázquez-Romero, Mårten S.G. Ahlquist, and Belen Martin-Matute J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.6b08350 • Publication Date (Web): 16 Sep 2016 Downloaded from http://pubs.acs.org on September 16, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of the American Chemical Society is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 8

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Base-Catalyzed Stereospecific Isomerization of ElectronDeficient Allylic Alcohols and Ethers through Ion-Pairing Samuel Martinez-Erro,a‡ Amparo Sanz-Marco,a‡ Antonio Bermejo Gómez,a Ana Vázquez-Romero,a Mårten S. G. Ahlquist,b and Belén Martín-Matutea* a

Department of Organic Chemistry, Stockholm University, Stockholm, SE-10691, Sweden. Division of Theoretical Chemistry & Biology, School of Biotechnology, KTH Royal Institute of Technology, Stockholm, SE-10691, Sweden. Allylic alcohols • Allylic ethers • Base catalysis • stereospecific • Ion pair b

Supporting Information

ABS TRACT: A mild base-catalyzed strategy for the isomerization of allylic alcohols and allylic ethers has been developed. Experimental and computational investigations indicate that transition metal catalysts are not required when basic additives are present. As in the case of using transition metals under basic conditions, the isomerization catalyzed solely by base also follows an sterospecific pathway. The reaction is initiated by a rate-limiting deprotonation. Formation of an intimate ion pair between an allylic anion and the conjugate acid of the base results in efficient transfer of chirality. Through this mechanism, stereochemical information contained in the allylic alcohols is transferred to the ketone products. The sterospecific isomerization is also applicable for the first time to allylic ethers, yielding synthetically valuable enantioenriched (up to 97% ee) enol ethers.

1. INTRODUCTION Allylic alcohols are readily accessible building blocks, and they are versatile synthons of carbonyl compounds via isomerization reactions. In the vast majority of instances, these are mediated by transition metal complexes.1,2 Basic additives, in catalytic or stoichiometric amounts, are also needed in numerous transition metal-mediated isomerizations. The base is postulated to deprotonate the substrate alcohol, which can subsequently coordinate to the metal center forming transition metal alkoxides, key intermediates in the proposed catalytic cycles. 2b-c,2h-i,3 With chiral metal catalysts, the synthesis of chiral ketones2a,g and aldehydes,2f-i with stereogenic centers at Cβ has been accomplished. An important related approach was reported recently by Cahard, Renaud and co-workers,4 who isomerized chiral γ-trifluoromethylated allylic alcohols sterospecifically using a non-chiral Ru complex together with Cs2CO3. Metal-free base-mediated protocols are much less common in the literature.5 These are highly dependent on the substrate structure (e.g., R1 = R3 = Ar), and require stoichiometric amounts, or excess, of the base. Metal-free chiral versions have not been reported. We have previously reported the synthesis of αhalocarbonyls from allylic alcohols catalyzed by Ir complexes.1c,6 As an on-going part of our investigations, we have now attempted to apply this methodology to electron-poor allylic alcohols, in particular γ-trifluoromethylated substrates. We have observed that these substrates isomerize efficiently under basic conditions, even at mild temperature, in the absence of the metal catalysts. Furthermore, the base-catalyzed isomerization achieves the same high levels of sterospecificity with

chiral allylic alcohols as those reported for the metal/base system.4 Herein, we report the first base-catalyzed sterospecific isomerization of allylic alcohols and also of allylic ethers (Scheme 1). The isomerizations are mediated by catalytic amount of the base 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD). Mechanistic experimental and theoretical investigations have been carried out allowing us to propose a mechanistic isomerization pathway accounting for the efficient transfer of chirality. Scheme 1. Sterospecific base-catalyzed isomerization of allylic alcohols and ethers R2 R3

OH

R2

TBD(cat)

O

R3 ∗

∗ R1

R1 N

TBD = R2 R3

OR 4 ∗

R1

TBD(cat)

R2 R3 ∗

OR 4

N H

N

R1

2. RESULTS AND DISCUSSION We started by investigating how the base affected the outcome of the isomerization of γ-trifluoromethylated allylic alcohol 1a (Table 1). Cs2CO3, which was also used in the isomerization of 1a in combination with [RuCl2(PPh3)3],4 can itself promote the isomerization at 60 °C (Table 1, entry 1). Ketone 2a was also

ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Table 1. Base-mediated isomerization of 1a

Base

OH

O

Toluene, F 3C 60 °C, 18 h

F 3C

rac-1a

N

N

DMAP

rac-2a

N N

DABCO

N

N N H

N DBU

N TBD

Entry

Base (equiv.)

Conv. (%)a

Yield (%)b

1

Cs2CO3 (1.2)

65

45

2

NaH (1.2)

25

20

c

t-BuOK (1.2)

>99

33

4

DMAP (1.2)

99

3

a

By 19F NMR spectroscopy. bWith an internal standard (I.S.). Decomposition products were formed. dUnder an atmosphere of air. c

obtained when NaH and t-BuOK were used, albeit in low (entries 2 and 3). The use of organic bases such as 4(dimethylamino)pyridine (DMAP) and 1,4diazabicyclo[2.2.2]octane (DABCO)5f did not promote the isomerization (entries 4 and 5). On the other hand, with the stronger7 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) or with guanidine base TBD conversions were achieved (entries 6 and 7), especially with TBD. Importantly, TBD was the only base that could be used in a catalytic amount to give a quantitative yield of 2a (entry 8, and Table S1). The isomerization works under at atmosphere of air, and after full conversion is achieved, a simple aqueous work-up to remove the basic catalyst gives the final products in excellent yields. The scope of the reaction (Scheme 2) under the optimized conditions (Table 1, entry 8) was then investigated. Variation of the electronic properties of the aryl group at R1 by introducing either electron-withdrawing or electron-donating groups gave quantitative yields (84-95% isolated) for all substrates. Notably, the reactions were much faster with electron-poor substituents (Scheme 2, 2b-c). The bulkier naphthyl group was also well tolerated, giving 2f in >99% yield. Excellent yields were also obtained when the aryl substituent at R1 was replaced by the electron-poor CF3 group (2g), although slightly harsher conditions were required. Significantly lower yields

Page 2 of 8

were achieved when R1 was replaced by CH3 or H (Scheme 2, 2h-i). The effect of varying the olefin substituents was next studied. Allylic alcohols bearing electron-withdrawing or electron-donating groups on aryl subtituents at R2 gave the corresponding products in high yields in all cases (Scheme 2, 2jm). Replacement of the aromatic group at R2 by H or CH3 was well tolerated, and alcohols 1o and 1n were isomerized in 9599% yields. Alcohols 1p and 1q, with longer fluorinated chains at R3, were also isomerized in good isolated yields (62% and 71%, respectively). The isomerization of (E)-1a was compared to that of its stereoisomer (Z)-1a. The latter smoothly transformed into 2a, but a higher temperature and a longer reaction time were needed. Interestingly, the fluorinated motif on the olefin can be replaced by other non-fluorinated groups; (E)-1r (R2 = H, R3 = Ph) isomerized in excellent yield. Even aromatic 1s-u with terminal double bonds isomerized successfully. Allylic alcohols (E)- and (Z)-1v, with non-fluorinated substituents on the double bond, one of them a methyl group, gave lower yields than those obtained with the related 1a. In contrast, with alcohols having two aryl substituents on the double bond (1w, (E)1x, (Z)-1x, and (E)-1y), excellent yields were obtained. It can therefore be concluded that the substitution pattern is limited at R1 to either aryl or electron-poor moieties such as the CF3 group). On the other hand, the R2 and R3 substituents can be varied to a larger extent; these may be electron-poor fluorinated groups, but importantly, also aryl groups with varied electron density, or even just hydrogen. It is important to highlight that TBD is able to isomerize not only fluorinated and electron poor alcohols, but also alcohols such as 1s-u, for which transition metal catalysts under basic conditions have been used.1,2 The base-catalyzed isomerization can also be applied to fluorinated allylic ethers 3a-3d (Scheme 2),8,9 giving access to synthetically valuable enol ethers.10 Even non-fluorinated terminal allylic ethers 3e-f could be isomerized. Enol ethers 4a4d were formed as single (Z)-isomers, and 4e and 4f as a 92:8 and 97:3 Z/E mixtures respectively (See SI). Unfortunately, other allylic substrates such as allyl silyl ethers and allylic esters did not isomerized under the conditions described here (See SI, Scheme S1). We then investigated whether the reaction is sterospecific (Table 2 and Table S2).4,11 The isomerization of (R)-1a (99% ee) took place in excellent yield and with an enantiospecificity (es; % of the product formed by the sterospecific pathway as opposed to the competing unselective pathway) of 91% (entry 1). The es could be improved to 95% by decreasing the base loading up to 2.5 mol%, although longer reaction times (18 h with 5 mol% of TBD and 72 h with 2.5 mol% of TBD) were needed (entries 2 and 3). Gratifyingly, at 80 ºC the es was maintained and the reaction was completed in 18 h with a 2.5 mol% of TBD (entry 4). Other solvents were screened to assess the effect on the es (Table 2, entries 5-8). Interestingly, all solvent tested not only gave decreased es, but also considerably lower yields (entries 5-8). This effect was more pronounced in solvents with higher polarity.

2 ACS Paragon Plus Environment

Page 3 of 8

Journal of the American Chemical Society Scheme 2. Scope

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Yield by 19F NMR spectroscopy (isolated in parenthesis). a80 °C. bTBD (0.2 equiv.), reflux. cBy NMR spectroscopy with an internal standard. dVolatile. eConversion.

ACS Paragon Plus Environment

3

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Table 2. Base-mediated sterospecific isomerization of 1a Ph F 3C

OH Ph

(R)-1a (99% ee)

Ph

TBDcat

60- 80 °C

Page 4 of 8

Table 3. Sterospecific isomerizations

O

F 3C

R2

Ph

(R)-2a

OR 4

R3

Entry

TBD (mol%)

Solvent

t (h) / T (°C)

2a (%)a

ee (%)

es (%)

1

10

Toluene

6 / 60

85

90

91

2

5

Toluene

18 / 60

84

93

94

3

2.5

Toluene

72 / 60

84

94

95

4

2.5

Toluene

18 / 80

82

94

95

5

2.5

1,4Dioxane

18 / 80

71

84

85

6

2.5

THF

18 / 80

30

74

75

7

2.5

CH2Cl2

18 / 80

28

80

81

8

2.5

CH3CN

18 / 80

27

60

61

a

Isolated yields.

The sterospecific isomerization was then tested for a variety of substrates (Table 3). As expected, the same es (91%) was obtained in the isomerization of (S)-1a (entry 2). (R)-1b and (R)-1c having electron-poor aryl groups at R1 gave the products with excellent es (96% and 98% respectively, entries 4 and 5). For alcohols with electron-rich R1, (R)-1d and (R)-1e (entries 6 and 7), longer reaction times were needed to achieve good yields with es of 94% and 89%, respectively. Naphthyl substituted (R)-1f gave similar results (entry 8). Lower yields were obtained in the isomerization of allylic alcohols (R)-1j and (R)-1k, with electron-poor aromatics at R2 (entries 9-10). In addition, whereas a p-Cl substituent did not significantly affect the es (entry 9), a somewhat less efficient transfer of chirality was achieved when R2 = p-CF3C6H4 (entry 10). On the other hand, exceptionally high es and good yields were obtained with alcohols (R)-1l and (R)-1m, with electrondonating aromatics at R2 (entries 11-12). Good transfer of chirality was also obtained with substrates bearing various fluorinated chains at R3 [(R)-1p and (R)-1q, entries 13-14]. Importantly, the ee of the products could be increased up to 99% by simple recrystallization. It must be also mentioned that the isomerizations are also sterospecific when bases such as Cs2CO3, NaH, t-BuOK and DBU were used (See SI, Table S1). The isomerization of chiral allylic ethers (R)-3a-c also proceeded remarkably well with 5 mol% of TBD at 80 °C (9799% es, entries 15-17). To the best of our knowledge, this is the first sterospecific isomerization of allylic ethers to be reported. The electronic influence of R1 was investigated further. A Hammett plot under non-competition conditions is shown in Figure 1. A positive ρ value of 1.87 was obtained, indicating that a partial negative charge (δ-) is formed in the ratedetermining transition state.12 These results, together with a large kinetic isotope effect (KIE) of ≥5.0, support a mechanism through a rate-limiting deprotonation of C1.13

R2

TBDcat

R1

R3

Toluene (R)-1 = H) or 80 °C (R)-3 (R 4 ≠ H) (R 4

Ph F 3C

R2

OH R1

R1

(R)-2

(R 4

OH

F 3C

R1 = (R)-1a C6H 5 (R)-1b p-BrC 6H 4 (R)-1c p-CF3C6H 4 (R)-1d p-MeC 6H 4 (R)-1e p-OMeC6H 4 (R)-1f 2-Naphthyl

or

= H)

Ph

R2 = p-ClC6H 4 p-CF3C6H 4 p-MeC 6H 4 p-OMeC6H 4 Ph

OR 4

R3

R1

(R)-4

(R 4

OH

R3

Ph

(R)-1j (R)-1k (R)-1l (R)-1m

R2

O

Ph R3 = (R)-1p C 2F 5 (R)-1q C3F 7

≠ H)

Ph

OMe

R3

Ph

R3 = (R)-3a CF3 (R)-3b C3F 7

OPh

F 3C

Ph

(R)-3c

Entry

ee 1/3 (%)

TBD (mol%)

t (h)

Yield 2/4 (%)a

ee 2/4 (%)b

es (%)

1

(R)-1a (99)

10

6

85

90

91

2

(S)-1a (-99)

10

6

92

-90

91

3

(R)-1a (99)

2.5

18

82

94 (99)

95

4

(R)-1b (98)

2.5

18

93

94 (99)

96

5

(R)-1c (97)

2.5

18

76

95 (99)

98

6

(R)-1d (98)

2.5

96

75

92 (99)

94

7

(R)-1e (98)

5

48

89

88 (99)

90

8

(R)-1f (98)

2.5

24

60

87

89

9

(R)-1j (97)

2.5

48

58

86

89

10

(R)-1k (97)

5

48

52

59

61

11

(R)-1l (97)

2.5

24

81

95

98

12

(R)-1m (98)

5

48

83

98

100

13

(R)-1p (99)

5

48

58

92

93

14

(R)-1q (99)

5

48

42

95

96

15

(R)-3a (98)

5

48

67

97

99

16

(R)-3b (98)

5

72

66

95

97

17

(R)-3c (98)

5

18

73

97

99

a

Isolated yields. bIn parentheses, ee after recrystallization.

4 ACS Paragon Plus Environment

Page 5 of 8

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

OH F 3C

3

2

TBD (0.1 equiv.) Toluene 60 °C

1

O F 3C

X

3

2

1

X

Figure 1. Hammett plot.

The isomerization of 1a-d1 (96% D) afforded 2a-d with only 68% of DTOT (20% at C2 and 48% C3 distributed in a mixture of 2a-d0 (32%), 2a-d1 (47%), 2a-d2 (17.6%) and 2a-d3 (3.4%), Scheme 3a and SI). In an attempt to understand the net low amount of D at C3, 1a was treated with partially deuterated TBD (N-d1-TBD, Scheme 3b). Notably, 6% of D was found at C3 in 2a, indicating that the H at C3 in Scheme 3a originates to some extent from the base. Importantly, in the isomerization of 1a-d1, the D content at C1 (i.e., 96%) remains constant in the starting alcohol as the reaction proceeds, thus the deprotonation step is irreversible. The isomerization of ether 3a-d1 (96% D, Scheme 3c) gave 4a-d1 with 78% D, which is only slightly lower than the expected 86% D. This might be due to traces of water in the reaction medium. From a crossover experiment

between 1a-d1 and 1n, with similar reaction rates (Scheme 3d), it can be concluded that the main reaction pathway involves an intramolecular 1,3-H/D shift, since only 4% of D was detected at C3 in 2n. To further test the proposed mechanism via deprotonation of the C1−H bond, we studied the isomerization with DFT (B3LYP-D3/aug-cc-pVTZ) (Figure 2). In agreement with the experimental studies, the deprotonation was found to be the rds, with an activation free energy of 19.6 kcal/mol (TSI), relative to the hydrogen bonded pre-reactive complex. After this first step, an intimate ion pair14 composed of the allylic anion and the protonated base is formed. The computational calculations suggest that the anion can be protonated either intramolecularly (TSIIa) by the hydroxyl group, or intermolecularly by the conjugate acid of TBD (TSIIb). Both pathways have similarly low activation energies (~4 kcal/mol). This explains the low deuterium content at C3 in the ketone product obtained upon isomerizing 1a-d1 (Scheme 3a). The theoretical studies also shed light onto the enantiospecificity of the reaction. The formation of the intimate ion pair prevents racemization of the allylic anion, which would result in inefficient transfer of chirality. Indeed, increasing the polarity of the solvent resulted in lower es values, due to insufficient ion pairing (vide supra, Table 2 and Table S2). Furthermore, the presence of a strong electron-withdrawing group on the phenyl ring at R2 (i.e. 2k) results in poorer transfer of chirality. This may be due to formation of a more stable allylic anion in this instance, what enables rotation with subsequent partial loss of stereochemical information. A similar mechanistic rationale has been proposed for a proton migration with chirality transfer mediated by DBU.15 The isomerization mechanism proposed here differs from that proposed for the Ru-catalyzed isomerization of trifluoromethylated allylic alcohols in the presence of base.4 In the latter, after a β-hydride elimination step yielding a ruthenium hydride species and an enone intermediate, it was proposed

Scheme 3. Deuterium labeling studies

5 ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

that a migratory insertion from a single face accounted for the efficient chirality transfer. Overall, this resulted in a formal Ru-catalyzed intramolecular suprafacial hydride shift. We have demonstrated here that for these particular electron-poor alcohols, the sterospecific isomerization can be simply mediated by a base through a deprotonation / protonation sequence that occurs at mild reaction conditions. Importantly, the basemediated reaction can account for the same levels of enantiospecificity. The best results during our investigations were obtained with TBD, which could be used in loadings as low as 2.5 mol%. Other bases such as Cs2CO3, NaH, of DBU also mediated the sterospecific isomerization of allylic alcohols, although in stoichiometric amounts. It is important to mention that higher yields are obtained with the Ru/Cs2CO34 system than when Cs2CO3 is used as the sole reagent, thus the Ru catalyst must play a specific role together with the base in the former case. Nevertheless, the non-innocent role of basic additives in the isomerization of electron-poor or aromatic allylic alcohols must be taken into consideration.

3. CONCLUSION We have developed the first metal-free base-catalyzed sterospecific isomerization of allylic alcohols. The method enables the transformation of a large variety of aromatic and electrondeficient allylic alcohols into carbonyl compounds under mild reaction conditions. We have also presented the first sterospecific isomerization of allylic ethers. This base-catalyzed metalfree method affords chiral enol ethers, which are also very

Page 6 of 8

valuable synthetic intermediates in organic synthesis. Experimental and computational investigations indicate that the transfer of the chirality only occurs upon formation of an intimate ion pair after the rate limiting deprotonation step. This study represents a unique example in which organic ionpairing enables transfer of chirality in readily available chiral organic allylic substrates using a simple non-chiral base.

4. EXPERIMENTA L SECTION General method for the isomerization of allylic alcohols and ethers by base. A pressure tube was charged with the corresponding allylic substrate (0.18 mmol, 1 equiv.), 1,5,7triazabicyclo[4.4.0]dec-5-ene (0.018 mmol, 2.5 mg, 0.1 equiv.) and 1.8 mL of toluene. The mixture was allowed to react at 60 °C overnight and the reaction was quenched with H2O (5 mL) and extracted with Et2O (3 x 5 mL). The solvent was removed under reduced pressure to yield the product. Further information about experimental details and characterization of new compounds can be found in the Supporting Information.

ASSOCIATED CONTENT Supporting Information Experimental details. Copies of 1H, 13C and 19F NMR spectra. Chromatograms of racemic and chiral products. The Supporting Information is available free of charge on the

Figure 2. Free energy profile: isomerization of 1a by TBD.

ACS Paragon Plus Environment

6

Page 7 of 8

Journal of the American Chemical Society

ACS Publications website.

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

AUTHOR INFORMATION Corresponding Author * Prof. B. Martín-Matute E-mail: [email protected]

Author Contributions All authors have given approval to the final version of the manuscript. ‡These authors contributed equally.

ACKNOWLEDGMENT This project was supported by the Swedish Research Council (VR), and the Knut and Alice Wallenberg Foundation. B.M.-M. was supported by VINNOVA through a VINNMER grant and A.V.-R. by the Wenner-Gren Foundation through a postdoctoral grant.

REFERENCES

(1) For reviews and accounts on transition metal-catalyzed allylic alcohol isomerizations, see: (a) Uma, R.; Crevisy, C.; Gree, R. Chem. Rev. 2003, 103, 27-52. (b) Cadierno, V.; Crochet, P.; Gimeno, J. Synlett. 2008, 1105-1124. (c) Ahlsten, N.; Bartoszewicz, A.; MartínMatute, B. Dalton Trans. 2012, 41, 1660-1670. (d) Lorenzo-Luis, P.; Romerosa, A.; Serrano-Ruiz, M. ACS Catal. 2012, 2, 1079-1086. (e) Cahard, D.; Gaillard, S.; Renaud, J. L. Tetrahedron Lett. 2015, 56, 6159-6169. (f) Li, H.; Mazet, C. Acc. Chem. Res. 2016, 49, 12321241. (2) Selected transition metal-catalyzed examples: (a) Ito, M.; Kitahara, S.; Ikariya, T. J. Am. Chem. Soc. 2005, 127, 6172-6173. (b) Martín-Matute, B.; Bogár, K.; Edin, M.; Kaynak, F. B.; Bäckvall, J.E. Chem. Eur. J. 2005, 11, 5832-5842. (c) Cadierno, V.; GarcíaGarrido, S. E.; Gimeno, J.; Varela-Álvarez, A.; Sordo, J. A. J. Am. Chem. Soc. 2006, 128, 1360-1370. (d) Voronova, K.; Purgel, M.; Udvardy, A.; Bényei, A. C.; Kathó, A.; Joó, F. Organometallics 2013, 32, 4391-4401. (e) Lastra-Barreira, B.; Díez, J.; Crochet, P. Green Chem. 2009, 11, 1681-1686. For asymmetric isomerizations of primary allylic alcohols, see: (f) Tanaka, K.; Qiao, S.; Tobisu, M.; Lo, M. M.-C.; Fu, G. C. J. Am. Chem. Soc. 2000, 122, 9870-9871. (g) Mantilli, L.; Gérard, D.; Torche, S.; Besnard, C.; Mazet, C. Angew. Chem. Int. Ed. 2009, 48, 5143-5147. (h) Wu, R.; Beauchamps, M. G.; Laquidara, J. M.; Sowa Jr, J. R. Angew. Chem. Int. Ed. 2012, 51, 2106-2110. (i) Arai, N.; Sato, K.; Azuma, K.; Ohkuma, T. Angew. Chem. Int. Ed. 2013, 52, 7500-7504. For asymmetric isomerizations of secondary cyclic allylic alcohols, see: (j) Kress, S.; Johnson, T.; Weisshar, F.; Lautens, M. ACS Catal. 2016, 6, 747-750. And reference 2a. (3) (a) Trost, B. M.; Kulawiec, R. J. J. Am. Chem. Soc. 1993, 115, 2027-2036. (b) Trost, B. M.; Indolese, A. F.; Müller, T. J. J.; Treptow, B. J. Am. Chem. Soc. 1995, 117, 615-623. (4) (a) Bizet, V.; Pannecoucke, X.; Renaud, J.-L.; Cahard, D. Angew. Chem. Int. Ed. 2012, 51, 6467-6470. (b) Bizet, V.; Pannecoucke, X.; Renaud, J.-L.; Cahard, D. J. Fluorine Chem. 2012, 152, 56-61. (5) (a) Dimmel, D. R.; Fu, W. Y.; Gharpure, S. B. J. Org. Chem. 1976, 41, 3092-3096. (b) Schmid, G. A.; Borschberg, H.-J. Helv. Chim. Acta. 2001, 84, 401-415. (c) Johnston, A. J. S.; McLaughlin, M. G.; Reid, J. P.; Cook, M. J. Org. Biomol. Chem. 2013, 11, 76627666. (d) Zheng, H.-X.; Xiao, Z.-F.; Yao, C.-Z.; Li, Q.-Q.; Ning, X.S.; Kang, Y.-B.; Tang, Y. Org. Lett. 2015, 17, 6102-6105. (e) Mondal, K.; Mondal, B.; Pan, S. C. J. Org. Chem. 2016, 81, 4835-4848. (6) (a) Ahlsten, N.; Martín-Matute, B. Chem. Commun. 2011, 47, 8331-8333. (b) Ahlsten, N.; Bermejo Gómez, A.; Martín-Matute, B. Angew. Chem. Int. Ed. 2013, 52, 6273-6276. (c) Bermejo Gómez, A.;

Erbing, E.; Batuecas, M.; Vázquez-Romero, A.; Martín-Matute, B. Chem. Eur. J. 2014, 20, 10703-10709, and references therein. (7) Superbases in catalytic reactions: Nacsa, E. D.; Lambert, T. H. J. Am. Chem. Soc. 2015, 137, 10246-10253. (8) Selected recent examples of transition-metal-catalyzed isomerization of allylic ethers: (a) Courchay, F. C.; Sworen, J. C.; Ghiviringa, I.; Abboud, K. A.; Wagener, K. B. Organometallics, 2006, 25, 60746086. (b) Varela-Álvarez, A.; Sordo, J. A.; Piedra, E.; Nebra, N.; Cadierno, V.; Gimeno, J. Chem. Eur. J. 2011, 17, 10583-10599. (c) Urbala, M. Applied Catalysis, A: General 2015, 505, 382-393. (9) Scarce examples of base-mediated isomerization of allylic ethers: (a) Su, C.; Williard, P. G. Org. Lett. 2010, 12, 5378-5381. (b) Reid, J. P.; McAdam, C. A.; Johnston, A. J. S.; Grayson, M. N.; Goodman, J. M.; Cook, M. J. J. Org. Chem. 2015, 80, 1472-1498. (9) Enol ethers as synthetic intermediates: (a) Modern Methods in Stereoselective Aldol Reactions, First Edition (Ed. R. Mahrwald), Wiley-VCH, 2013. (b) Kusama, H.; Ebisawa, M.; Funami, H.; Iwasawa, N. J. Am. Chem. Soc. 2009, 131, 16352-16353. (b) Vitale, A.; Bongiovanni, R.; Ameduri, B. Chem. Rev. 2015, 115, 8835-8866. (11) Synthesis of chiral allylic alcohols and synthetic applications: Lumbroso, A.; Cooke, M. L.; Breit, B. Angew. Chem. Int. Ed. 2013, 52, 1890-1932. (12) (a) Hammett, L. P. J. Am. Chem. Soc. 1937, 59, 96-103. (b) Lupp, D.; Christensen, N. J.; Fristrup, P. Understanding Organometallic Reaction Mechanisms and Catalysis: Computational and Experimental Tools, First Edition (Ed. V. P. Ananikov ), Wiley-VCH, 2015, pp 187-215. (13) Simmons, E. M.; Hartwig, J. F. Angew. Chem. Int. Ed. 2012, 51, 3066-3072. (14) Ion-pairing in catalytic processes: (a) Rueping, M.; Uria, U.; Lin, M.-Y.; Atodiresei, I. J. Am. Chem. Soc. 2011, 133, 3732-3735. (b) Brak, K.; Jacobsen, E. N. Angew. Chem. Int. Ed. 2013, 52, 534561. (c) Mahlau, M.; List, B. Angew Chem., Int. Ed. 2013, 52, 518533. (d) Xue, H.; Jiang, D.; Jiang, H.; Wee Kee, C.; Hirao, H.; Nishimura, T.; Wah Wong, M.; Tan, C.-H. J. Org. Chem. 2015, 80, 57455752. (15) For a related DBU-catalyzed stereospecific isomerization through proton transfer, see: Dabrowski, J. A.; Haeffner, F.; Hoveyda, A. H. Angew. Chem. Int. Ed. 2013, 52, 7694-7699.



7 ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 8

Metal-Free Enantiospecific Isomerization Allylic Alcohols R2 R3

Allylic Ethers

OH H R1

N N H

R3

O

R2 R1

R3

O

2

O

R3

N H

2 H R

R2

OR4 R1

H

HR R3

R4 H Ph R4 Ph

- Catalytic - Mild conditions - Broad scope

ACS Paragon Plus Environment

8